Certain Metal Ions Are Inhibitors of Cytochrome b6f Complex 'Rieske

The Raised Midpoint Potential of the [2Fe2S] Cluster of Cytochrome bc1 Is Mediated by ... Alternative Rieske Iron-Sulfur Subunits and Small Polypeptid...
0 downloads 0 Views 142KB Size
4070

Biochemistry 2002, 41, 4070-4079

Certain Metal Ions Are Inhibitors of Cytochrome b6f Complex ‘Rieske’ Iron-Sulfur Protein Domain Movements† Arthur G. Roberts,‡ Michael K. Bowman,‡,§ and David M. Kramer*,‡ Institute of Biological Chemistry, Washington State UniVersity, 289 Clark Hall, Pullman, Washington 99164-6340, and WR Wiley EnVironmental Molecular Sciences Laboratory, Pacific Northwest National Laboratory, Richland, Washington 99352-0999 ReceiVed NoVember 29, 2001; ReVised Manuscript ReceiVed January 31, 2002 ABSTRACT: Many current models of the Q cycle for the cytochrome (cyt) b6f and the cyt bc1 complexes incorporate ‘Rieske’ iron-sulfur protein (ISP) domain movements to gate electron transfer and to ensure high yields of proton shuttling. It was previously proposed that copper ions, which bind at a site distant from the quinol oxidase (Qo) site, inhibit plastoquinol (PQH2) binding by restraining the hydrophilic head domain of the ISP [Rao B. K., S., Tyryshkin, A. M., Roberts, A. G., Bowman, M. K., and Kramer, D. M. (1999) Biochemistry 38, 3285-3296]. The present work presents evidence that this is indeed the case for both copper ions and Zn2+, which appear to inhibit by similar mechanisms. Electron paramagnetic resonance (EPR) spectra show that Cu2+ and Zn2+ binding to the cyt b6f complex displaces the Qo site inhibitor 2,5-dibromo-3-methyl-6-isopropylbenzoquinone (DBMIB). At high concentrations, both DBMIB and Cu2+ or Zn2+ can bind simultaneously, altering the Rieske 2Fe2S cluster and Cu2+ EPR spectra, suggesting perturbations in their respective binding sites. Both Zn2+ and Cu1+ altered the orientations of the Rieske 2Fe2S cluster with respect to the membrane plane, but had no effect on that of the cyt b6 hemes. Cu2+ was found to change the orientation of the cyt f heme plane, consistent with binding on the cyt f protein. Within conservative constraints, the data suggest that the ISP is shifted into a position intermediate between the ISPC position, when the Qo site is unoccupied, and the ISPB position, when the Qo site is occupied by inhibitors such as DBMIB or stigmatellin. These results support the role of ISP domain movements in Qo site catalysis.

The cytochrome (cyt)1 b6f complex, which is analogous both in structure and in function to the mitochondrial and bacterial cyt bc1 and bc-type complexes, plays a central role in photosynthetic electron transport in cyanobacteria and plant chloroplasts (for reviews, see 1-9). It serves three main functions: (1) It transfers electrons from plastohydroquinone, or plastoquinol (PQH2), to a soluble carrier, plastocyanin (PC) or a c-type cyt, which delivers them to photosystem (PS) I. (2) It uses the free energy available in the electrontransfer reactions to shuttle protons from the stroma (n-side) to the lumen (p-side), contributing to the establishment of proton motive force (pmf), which drives the synthesis of ATP. † This work was supported by U.S. Department of Energy Grant DE-FG03-98ER20299, by a Frasch Foundation award (D.M.K.), and by NIGMS Grant GM61904 (M.K.B.). The WR Wiley Environmental Molecular Sciences Laboratory is a national scientific user facility sponsored by the Department of Energy’s Office of Biological and Environmental Research and located at Pacific Northwest National Laboratory. * Corresponding author. Phone: ++(509) 335-4964; Fax: ++(509) 335-7643; Email: [email protected]. ‡ Washington State University. § Pacific Northwest National Laboratory. 1 Abbreviations: cyt, cytochrome; cyt b , low-potential b cytochrome; L cyt bH, high-potential b cytochrome; DBMIB, 2,5-dibromo-3-methyl6-isopropylbenzoquinone; DMSO, dimethyl sulfoxide; DNP-INT, 2-iodo-6-isopropyl-3-methyl-2′,4,4′-trinitrodiphenyl ether; EPR, electron paramagnetic resonance; HEPES, N-(2-hydroxyethyl)piperazine-N′-2ethanesulfonic acid; ISP, iron-sulfur protein; pmf, proton motive force; PC, plastocyanin; PQ, plastoquinone; PQH2, plastoquinol; PS, photosystem; Qi, quinol reductase; Qo, quinol oxidase; UHDBT, 5-n-undecyl6-hydroxy-4,7-dioxobenzothiazole.

(3) It acts as a redox sensor, initiating antenna state transitions when the plastoquinone (PQ) pool becomes predominantly reduced (10-13). The cyt b6f complex is comprised of four major protein subunits: the cyt f protein, the ‘Rieske’ iron-sulfur protein (ISP), the cyt b6 protein, and subunit IV (14). In the cyt bc1 complex, the latter two are fused into a single analogous protein, known as the cyt b subunit. Both the cyt b6 and the cyt b subunits house one low-potential cyt b heme (cyt bL) and one high-potential cyt b heme (cyt bH). The cyt f subunit contains an unusual c-type heme with one axial ligand supplied by the protein N-terminal amino group (15) and is functionally analogous to the cyt c1 protein in the mitochondrial cyt bc1 complex. The ISP houses a ‘Rieske’-type ironsulfur cluster (2Fe2S), which is coordinated by two histidines and two cysteines. Several smaller subunits, chlorophyll, and carotenoid molecules have also been found in the cyt b6f complex, but their functions are not well understood (1, 1618). The cyt b6f complex also contains two PQ/PQH2 binding sites which function during normal turnover to oxidize PQH2 at the quinol oxidase (Qo) site and reduce PQ at the quinol reductase (Qi) site. The Qi site is formed by the cyt b6 protein, and subunit IV, toward the n-side of the membrane (i.e., stroma) (reviewed in 1, 3, 4, 14). The analogous site in the cyt bc1 complexes is found in the cyt b subunit. The Qo site of the cyt b6f complex is at the interface between cyt b6 and the Rieske ISP and adjacent to the p-side of the membrane

10.1021/bi015996k CCC: $22.00 © 2002 American Chemical Society Published on Web 03/03/2002

Inhibitory Metals Alter Cytochrome b6f ISP Orientation (i.e., thylakoid lumen). Several inhibitors bind strongly to this site, including stigmatellin, 5-n-undecyl-6-hydroxy-4,7dioxobenzothiazole (UHDBT), 2-iodo-6-isopropyl-3-methyl2′,4,4′-trinitrodiphenyl ether (DNP-INT), and 2,5-dibromo3-methyl-6-isopropylbenzoquinone (DBMIB) (2-4, 19). Electron transfer through the complex occurs through a process known as the Q-cycle, first proposed by Mitchell (20), and later modified by several groups (see reviews 2126). Electron transfer from the Qo site quinol is bifurcated, so that one electron is transferred to the ‘high potential chain’, consisting of the 2Fe2S cluster and cyt f, while the other is forced through the ‘low potential chain’ consisting of cyt bL and cyt bH to a quinone at the Qi site. This results in the release of two protons from the Qo site quinol into the chloroplast lumen, contributing to pmf that drives ATP synthesis. The yield of the bifurcated electron transfer and proton pumping in the higher plant cyt b6f complex is high both in vitro (27) and in vivo (28), implying that reactions which bypass the Q-cycle are minimized. Since transferring both quinol electrons to the high-potential chain would be heavily favored by thermodynamics (see discussions in 27, 29), it has long been proposed that a ‘catalytic switch’ gates electron transfer at the Qo site, to force electron flow down the two pathways (30, 31). The recent structural and functional advances in our understanding of the cyt bc1 complex offer an intriguing explanation for such a catalytic switch. Different crystal forms and crystals made in the absence and presence of Qo site inhibitors showed that the hydrophilic ‘head’ domain of the ISP occupies several different positions. One position, termed ISPB, places the ISP close to the distal niche of the Qo site; another, termed ISPC, places the ISP close to cyt c1; still others place the ISP in intermediate positions (2, 32-35). Molecular dynamics simulations of the ISP in these structures showed that the ISP likely rotates freely around a central ‘pivot’ by 57°-63° (36). This pivoting action of the ISP is thought to prevent the transfer of both electrons from the quinol to the high-potential chain, and is consistent with the observed electron-transfer rates (33, 36). Electron transfer between quinol bound at the Qo site and the nearby 2Fe2S cluster, while in the ISPB configuration, should produce a reduced 2Fe2S cluster and a semiquinone at the Qo site, locking it into this configuration (as suggested in 33, 3740). As long as the ISP is prevented from adopting the ISPC position, reoxidation of the 2Fe2S cluster would be hindered by the large cluster to cyt c1 heme distance, leaving cyt bL as the only available oxidant for the semiquinone. Conversely, when in the ISPC position, the 2Fe2S cluster is too far from the Qo site to allow oxidation of the QH2 or the semiquinone. Thus, coordinated motion of the ISP head domain between the ISPB and ISPC positions is key to the operation of the Q-cycle. We note that, although the cyt b6f complexes contain cyt f instead of cyt c1, we retained the term ISPC to be consistent with the cyt bc1 literature. It has been known for many years that Zn2+ and other divalent metal ions inhibit cyt bc1 complexes (41-45). In previous work, we showed that Cu2+ inhibits the cyt b6f complex (46). There are important differences between the mechanisms of metal inhibition of the cyt bc1 and cyt b6f complexes. Link et al. (41) found that Zn2+ inhibition of the cyt bc1 complex was noncompetitive with either quinol or cyt c substrates. This, together with the pH-dependence of

Biochemistry, Vol. 41, No. 12, 2002 4071 the Ki, suggested that Zn2+ may act by blocking proton release from the Qo site. Since then, Berry et al. (47) showed that avian mitochondrial cyt bc1 complex contains at least two Zn2+ binding sites: one near the Qo site, which was hypothesized to cause inhibition, and the other close to the interface between cyt b and cyt c1. In contrast, our work has shown that inhibition of the spinach cyt b6f complex by Cu2+ (46) and Zn2+ (see below) was competitive with the substrate PQH2, but noncompetitive with PC or horse heart cyt c, suggesting substantially different mechanisms for inhibition. One great advantage of Cu2+ as an inhibitor is that it is paramagnetic, allowing us to use pulsed-EPR techniques to probe its binding site. With these techniques, we showed that Cu2+ binds to a histidine residue (46). A structural model of the cyt b6f complex (cf. Figure 10 of 46), based on homology modeling to the X-ray crystal structures of the mitochondrial cyt bc1 complex (33) as well as the structures of the cyt b6f complex ISP (48, 49) and cyt f protein (3, 15), suggested that the Cu2+ binding site was probably on cyt f or the ISP. Because these residues lie distant from the Qo site, we concluded that Cu2+ binding displaces PQH2, probably via longer-range conformational changes that are transmitted through the ISP (46). Thus, we proposed that metal ions inhibit the turnover of the cyt b6f complex by restraining normal ISP domain movements. MATERIALS AND METHODS Thylakoid Preparation. Spinach thylakoids were prepared essentially as described in (27) with one modification. In the final stage, the thylakoid pellet was resuspended in HCL buffer [i.e., 30 mM N-(2-hydroxyethyl)piperazine-N′-2ethanesulfonic acid (HEPES, pH 7.6), 0.5% cholic acid, 0.1% R-lecithin] with 10% (NH4)2SO4 and 5-10% dimethyl sulfoxide (DMSO) to a concentration of greater than 4 mg mL-1 chlorophyll and stored at -80 °C. The chlorophyll content of the thylakoids was determined by the method of Arnon (50). The choice of HEPES buffer was critical since many other buffers chelate Cu2+ and other metals, preventing strong interaction with the cyt b6f complex (51). Purification of the Cyt b6f Complex. Cytochrome b6f complex was isolated from spinach thylakoids essentially as described by Hurt and Hauska (52, 53) with some modifications. Thylakoids at 3 mg mL-1 chlorophyll were solubilized with HCL buffer containing 40-60 mM n-octyl glucoside and 10% (NH4)2SO4 and incubated 30 min on ice in the dark. The precise concentration of detergent was determined for each thylakoid batch as the concentration required for the release of ∼90% cyt b6f complex from the membranes (53, 54). Detergent-solubilized suspensions were centrifuged at 141000g for 30 min at 4 °C. Then this was diluted with HCL buffer containing 40% (NH4)2SO4 to the supernatant to give a final concentration of (NH4)2SO4 of 25%. Ammonium sulfate fractionation was performed with most of the cyt b6f precipitating in the 45-55% fraction as has been previously described (53). The pellet was resuspended in HCL buffer with 30 mM n-octyl glucoside. This was dialyzed for 1 h against HCL buffer containing 10 mM n-octyl glucoside to remove the (NH4)2SO4. The detergent n-octyl glucoside was added slowly to the dialyzed samples to a final concentration of 30 mM, and the resulting suspension was placed onto a sucrose density gradient, 7-30%, which was prepared by freeze-thaw technique using HCL buffer with 30 mM

4072 Biochemistry, Vol. 41, No. 12, 2002 n-octyl glucoside and was centrifuged at 141000g for 1216 h at 4 °C. After centrifugation, the resulting brown band in the sucrose density gradient was assayed for activity and cyt f concentration as previously described in (46). The brown band was then dialyzed against HCL buffer with 10 mM n-octyl glucoside for 1 h to remove the sucrose. Then the n-octyl glucoside was added to the dialyzed protein until the final concentration was 30 mM. The suspension was then concentrated to 10-30 µM cyt f, using a 100 kDa cutoff Amicon Centricon concentrator (Millipore, Inc., Bedford, MA), and then stored with 20% glycerol at -80 °C. The activity of the complex was measured as described in (46) with turnover of ∼30 s-1 at 100 µM cyt c. Preparation of Oriented Cyt b6f Complex. Partially ordered samples were obtained from purified cyt b6f complex as reported previously (37, 55-58). A stock suspension of ∼30 µM cyt b6f complex was dissolved in 25 mL of 30 mM HEPES, to a final concentration of approximately 0.4 µM cyt b6f complex. This was centrifuged for 14 h at 141000g and 4 °C. The pellet was then dissolved in 25 mL of distilled, deionized water and centrifuged for 30 min at 141000g and 4 °C. Inhibitors, including 10 µM Cu2+ or Zn2+, were added at this step. The resulting pellet was resuspended with a small amount (∼50-200 µL) of deionized, distilled water. This was applied to strips of Mylar (type A, Dupont Co., Wilmington, DE) with a small paintbrush and allowed to dry for 3 days under argon above 80% w/v ZnCl2 to maintain humidity. To reduce the 2Fe2S cluster or Cu2+, while maintaining the cyt b hemes oxidized, the strips were dipped momentarily in a 10 mM solution of sodium ascorbate buffered with 30 mM HEPES (pH 7.6). To obtain oxidized samples, where the cyt b and cyt f were oxidized, strips were briefly dipped in 5 mM sodium ferricyanide buffered with 30 mM HEPES (pH 7.6). After dipping, strips were allowed to dry under Ar(g) for several minutes. Several strips were then stacked and placed in an EPR tube and quickly stored at -80 °C. EPR Spectroscopy. Electron paramagnetic spectra were obtained using a Bruker 200tt EPR spectrometer (Bruker Instruments, Billerica, MA), using a GFS-300 transfer tube, ESR-900 helium cryostat with model ITC4 temperature controller (Oxford Instruments, Oxford, U.K.). The data were acquired using a CIO-DAS1401/12 digital I/O board (Computer Boards, Inc. Mansfield, MA) connected to a Pentium PC computer and software written in Microsoft Visual Basic 6.0 (Microsoft, Redmond, WA). The EPR acquisition parameters are provided in the figure legends. The orientation of the Mylar sheets within the magnetic field was adjusted and measured using a Varian E-229 goniometer (Varian, Inc., Palo Alto, CA) with an accuracy of about (1°. EPR Spectral Simulations of Oriented Samples. The EPR spectra of the oriented membrane samples were simulated in a two-step procedure, using Mathematica Version 4.1 (Wolfram Research, Champaign, IL) and standard equations (59). First, the EPR frequency, Ω, of the ISP or other paramagnetic center was scaled with respect to its values, ω, at gmax and gmin as

Y(Ω) )

Ω - ω(gmin) ω(gmax) - ω(gmin)

so that in the absence of inhomogeneous broadening of the EPR spectrum, 0 e Y e 1. A second scaled parameter, f )

Roberts et al. Y(gmid), characterizes the rhombicity, or deviation from axial symmetry of the g-tensor. When the magnetic field makes angles of θ and φ with respect to the gmax and gmin axes of the ISP, the EPR frequency, Y, is Υ(z,φ) ) f*(1 - z2)*sin(φ)2 + z2 where z ) cos(θ). Assuming isotropic inhomogeneous broadening of the EPR transition described by a Gaussian function with width proportional to dν, the natural logarithm of the un-normalized EPR intensity at the scaled frequency, ν, of an ISP at an arbitrary orientation is lnEPR(z,φ,ν) ) -[Y(z,φ) - ν/dν]2. The logarithm of the intensity is a much more smoothly varying function than the EPR intensity itself and provided much faster convergence in constructing an interpolating function for lnEPR within the cube defined over 0 e z e 1, 0 e φ eπ/2, and 0 e ν e 1 using Mathematica. This interpolating function then allows rapid integration over z and φ for calculation of the EPR spectrum of an arbitrary range of orientations. If the ISP adopts a favored orientation, then the 2Fe2S g-tensor axes with respect to the membrane will be defined by a set of Euler angles, (0,ξ,ψ). The preferred axis will deviate from the membrane perpendicular by an angle θ′ due to disorder or mosaic spread with a relative probability exp[-(θ′/m)2] where m parametrizes the mosaic spread. In the spectrometer, the magnetic field is at an angle R with respect to the membrane perpendicular. Thus, the magnetic field is related to the 2Fe2S cluster g-tensor, and thus the ISP, by the Euler angles (0,R,β). The angle β takes on all possible values because the ISP molecules are aligned only with respect to one direction. The probability of finding an ISP oriented at (z,φ) with respect to the magnetic field in a membrane oriented with an angle R between its perpendicular and the magnetic field is given by

[(

P(z,φ,R) ) ∫0 exp 2π

)]

arccos(F[β]) m

2

‚dβ

where

F(β) ) cos(R)‚[z‚cos(ξ) + x1 - z2‚cos(φ - ψ)‚sin(ξ)] + sin(R)‚{cos(β)‚[x1 - z2‚cos(φ - ψ)‚cos(ξ) z‚sin(ξ)] + x1 - z2‚sin(φ - ψ)‚sin(β)} and the full symmetry with respect to z and φ was used to fold the unit sphere into a single quadrant. Another interpolating function in Mathematica was fit to P(z,φ,R), and the intensity at any point in the EPR spectrum for a membrane at angle R is then readily calculated from the two interpolating functions as

I(ν,R) ) ∫∫P(z,φ,R)‚exp[lnEPR(z,φ,ν)] dzdφ For the polar plots, experimental data were compared with I(1,R) and I(0,R) at gmax and gmin, respectively, and to 2*I(f,R) - I(f+0.02, R) - I(f-0.02,R), which is proportional to the finite-difference approximation to the second derivative, at gmid. Each simulated polar plot was normalized so that the maximum value was unity for comparison to data. Use of a Gaussian function for the mosaic spread and an isotropic line width represent major limitations of these simulations. The mosaic spread in orientations has contributions from less than perfect orientation of the proteins within the membranes, including defects in the membrane such as

Inhibitory Metals Alter Cytochrome b6f ISP Orientation

FIGURE 1: Effect of Qo site inhibitors on cyt b6f complex bound Cu2+ EPR signal. EPR spectra of Cu2+ bound to 30 µM cyt b6f complex in (A) uninhibited, (B) 1 equiv (30 µM) of DBMIB, and (C) 1 equiv (30 µM) of stigmatellin. Arrows in the figure emphasize differences between the g| hyperfine peaks. EPR spectrometer settings were: microwave frequency ) ∼9.429 GHz; microwave power ) 6.32 mW; gain ) 4.0 × 105; time constant ) 1000 ms; modulation amplitude ) 1.6 mT; center field ) 370 mT; sweep width ) 200 mT; sweep time ) 2 min; averages ) 32; temperature ) 70 K.

holes, folds, edges, and inclusions and also irregularities in the surface of the Mylar sheets. It is unlikely that the actual distribution is Gaussian; however, given the relatively large values of m needed to simulate the rotational plots, the actual distribution function is likely to be broad and featureless. The EPR spectrum of the ISP has a large amount of g-strain that is not included in the line width model. To do so would make the simulations much slower without improving the results given the level of approximations in the mosaic spread function. To fit the data, the simulations were superimposed on the experimental data at 1-5° angle increments with a mosaic spread between 0.250*π and 0.350*π. The closest fits were determined by eye to have a mosaic spread of 0.275*π and angles that are described in the text. RESULTS Effects of Qo Site Occupancy on Cu2+ Ligand Geometry. Figure 1 shows EPR spectra of Cu2+ bound to the cyt b6f complex (unoriented samples) in the absence or presence of the Qo site inhibitors DBMIB and stigmatellin. In the absence of inhibitors, the Cu2+ EPR spectrum was essentially as previously observed (46), and typical of type II Cu2+ spectra (for reviews, see 60-64). The g| hyperfine line shapes indicate that two Cu2+ forms contribute to the observed spectra. The relative amounts of each form depend on the sample pH (Rao B. K., S., Tyryshkin, A. M., Kramer, D. M., and Bowman, M. K., unpublished data). These samples were prepared without ferricyanide to avoid overlap of its broad EPR signal at g ) ∼2.7 (60, 65). The Cu2+ g⊥ transition was found at g ) 2.04 with a line width of 14 mT, both in the absence and in the presence of DBMIB or stigmatellin. In the absence of Qo site inhibitors (Figure 1, trace A), the g| transition was found at g ) 2.22 with average Cu2+ hyperfine splittings of A| ) 2.21 × 10-2 cm-1. In the presence of DBMIB (Figure 1, trace B), g| was shifted to g ) ∼2.225 with A| ) 2.04 × 10-2 cm-1. On the other hand,

Biochemistry, Vol. 41, No. 12, 2002 4073 in the presence of stigmatellin (Figure 1, trace C), g| was shifted to g ) ∼2.235 with A| ) 1.98 × 10-2 cm-1. Since the g| region is sensitive to the ligand environment of the Cu2+ (60, 63, 66-68), we concluded that introducing inhibitors into the Qo site induced significant changes to the Cu2+ binding site. In samples in Figure 1, the 2Fe2S EPR signal showed only a small gy signal at g ) 1.89, indicating that the cluster was predominantly oxidized. In any case, samples prepared with ferricyanide or with a slight excess of ascorbate to reduce the ISP but not Cu2+ showed nearly identical Cu2+ EPR spectra, albeit with different contributions in the Cu2+ g⊥ region from the reduced 2Fe2S cluster (data not shown). From this, we concluded that the redox state of the 2Fe2S cluster did not have large effects on the ligand geometry of bound Cu2+. Zn2+ as an Inhibitor of the Cyt b6f Complex. As demonstrated above, the EPR-sensitivity of Cu2+ makes it a sensitive probe of its structural environment. However, for other experiments, Cu2+ can be a less than an ideal inhibitor. The free (69) and cyt b6f complex-bound (46) Cu2+/Cu1+ couples are readily reduced by reductants such as ascorbate. Cu1+ reacts rapidly with O2 to produce superoxide (see 70, 71), which can damage enzymes. In addition, the EPR spectrum of Cu2+ may overlap with other paramagnetic centers of interest. For these reasons, we also studied the effects of Zn2+, which is not redox-active within physiological conditions and does not reduce O2 (see 72). Kinetic assays showed that Zn2+ is competitive with PQH2 and noncompetitive with cyt c, as we found earlier for Cu2+ but with KI values somewhat higher than Cu2+ (∼4 µM at 1 µM decyl PQH2, A. G. Roberts and D. M. Kramer, in preparation). Although the inhibitory binding sites may differ, we propose that Zn2+ inhibits by the same general mechanism as copper. Because Zn2+ is diamagnetic, we were not able to use EPR to probe its binding site on the cyt b6f complex. Effects of Copper and Zinc on the Rieske 2Fe2S EPR Spectrum in the Absence of Qo Site Occupants. In the absence of Qo site inhibitors, addition of inhibitory concentrations of Cu1+, Cu2+, and Zn2+ had no apparent effects on the EPR spectrum of the reduced 2Fe2S cluster (data not shown). Such shifts would imply strong or direct interaction with the 2Fe2S cluster (see reviews 37, 73, 74). These results are consistent with previous studies of copper, which found that the metal was bound far from the Qo site and the 2Fe2S cluster (46). Effects of Cu2+ and Zn2+ on the DBMIB-Induced Shifts of the Rieske 2Fe2S EPR Spectrum. Figure 2 shows EPR spectral changes of the 2Fe2S cluster upon addition of metals and the Qo site inhibitor DBMIB. Addition of 2 equiv of DBMIB per complex was sufficient to induce the large, wellknown changes in the EPR spectrum (Figure 2, trace A) with the appearance and increase of the g ) 1.94 signal and the decrease in the g ) 1.89 signal, as has been previously demonstrated (39, 75, 76). The concentration of ascorbate was kept low to prevent reduction of free DBMIB [Em(pH 8.0) ) ∼60 mV, (77)] to its quinol form, which binds weakly to the complex (39). In the presence of DBMIB, the gz transition around g ) 2.03 showed distinct heterogeneity, expressed as an apparent second peak near g ) 1.99, which was previously attributed to a free radical signal (39). When Cu2+ and Zn2+ were added to DBMIB-treated complex, relatively large changes in the 2Fe2S spectrum were

4074 Biochemistry, Vol. 41, No. 12, 2002

Roberts et al.

FIGURE 2: Effect of metals on the DBMIB-shifted “Rieske” 2Fe2S EPR spectrum. EPR spectra of the 2Fe2S cluster (25 µM cyt b6f complex) with 2 equiv (50 µM) of DBMIB and (A) no metals, (B) 2 equiv (50 µM) of Cu2+, or (C) 4 equiv (100 µM) of Zn2+. Trace 2B (dotted line) shows the spectrum of Cu2+-bound cyt b6f complex spectra without reductants, where the 2Fe2S cluster was oxidized, illustrating the overlap of Cu2+ and 2Fe2S EPR spectra. EPR spectrometer settings were the same as in Figure 1 except the temperature was 20 K.

observed (Figure 2, traces B and C). Upon addition of equimolar concentrations of Cu2+ and DBMIB to the complex, the g ) 1.89 transition showed an increase in line width to 5.5 mT from 4.0 mT and shifted upfield by 0.5 mT to g ) ∼1.88. The amplitude of the transition at g ) ∼1.88 increased with respect to the g ) 1.94 transition in these samples, indicating that a fraction of the DBMIB was displaced (see 78) or that the ISP was no longer interacting strongly with the DBMIB. Unfortunately, the significant signal overlap from the Cu2+ EPR spectrum (see Figure 2, trace B, dotted line) prevented a thorough analysis of changes in the 2Fe2S gz ) 2.03 transition or the g ) 1.94 transition. Addition of a 4-fold excess of Zn2+ caused similar shifts in the 2Fe2S spectrum of DBMIB-treated complex (Figure 2, trace C), but in this case without introducing interfering EPR signals. The amplitude of the transition at g ) ∼1.88 also increased with concomitant decrease in the g ) 1.94 transition as was seen in the Cu2+-inhibited samples. Furthermore, the g ) ∼1.88 transition line width increased to ∼5.5 mT, similar to the Cu2+-inhibited samples. This was accompanied by a decrease in the g ) 1.94 signal that is characteristic of the displacement of DBMIB (78). Furthermore, the gz transition at g ) 2.03 appeared to shift to g ) 2.01. Additions of higher concentrations of Cu2+ or Zn2+ led to nonspecific binding (46) and precipitation of the metal (data not shown, see also 79) and were thus not useful for these studies. Similar experiments were also performed with Cu1+, but they were problematic because of the high concentrations of ascorbate that were used to keep the Cu1+ reduced also reduced DBMIB to its quinol form. Orientation of the Redox Carriers of the Cyt b6f Complex Estimated by EPR Spectroscopy of Mylar-Oriented Cyt b6f Complex. Figure 3A shows 2Fe2S EPR spectra of oriented

FIGURE 3: Orientations of the “Rieske” 2Fe2S cluster in the presence of metals. (A) EPR spectra of the “Rieske” 2Fe2S cluster in partially oriented samples of cyt b6f complex in the absence of metals with magnetic field at 0° (solid line) and 90° (dashed line) with respect to the membrane plane. The gx, gy, and gz transitions are found at 1.75, 1.89, and 2.03, respectively. Plots B-J show the angular dependence of the three EPR transitions on the angle of the magnetic field with respect to the membrane plane. The experimental data are shown as points, and the simulations (see Materials and Methods) are shown as solid lines superimposed on the experimental data. Data were taken in the absence of metals (plots B-D) and in the presence of Cu1+ (plots E-G) or Zn2+ (plots H-J). The orientation dependencies of gx are shown in plots B, E, and H, those for gy are shown in plots C, F, and I, and those for gz are shown in plots D, G, and J. The orientation of the membrane plane with respect to the plots is shown at the bottom of the figure. For reliable determinations of the orientation dependence of the gx transition, the amplitude of the transition was measured from a horizontal baseline drawn from 425 to 430 mT. EPR spectrometer settings were as in Figure 2, and a goniometer was used to orient the samples (see Materials and Methods).

cyt b6f complex in the absence of inhibitors with the magnetic field oriented at 0° (solid line) and 90° (dotted line) with respect to the membrane plane. There was significant anisotropy of all three, gx ) 1.75, gy ) 1.89, and gz ) 2.03, transitions. In Figure 3B-D, the magnetic field-orientation dependence of these transitions is plotted with respect to the membrane plane. The g-axis orientations are similar to those reported earlier (37, 39). Production of Mylar-oriented cyt b6f complex with Cu2+ and reduced (paramagnetic) 2Fe2S was not reproducible because of the reduction of Cu2+ or the oxidation of the

Inhibitory Metals Alter Cytochrome b6f ISP Orientation 2Fe2S cluster. Therefore, assays were performed with excess ascorbate, where Cu2+, based on EPR analysis, was probably reduced to Cu1+ (80). The orientation dependence of the EPR transitions of samples containing at least some Cu2+ as determined by EPR analysis gave virtually identical results (data not shown). In addition, kinetic assays with Cu1+ suggested qualitatively that Cu1+ also acts competitively with PQH2 and noncompetitively with the mobile redox carriers, PC and cyt c, although it is complicated because Cu1+ can subsequently reduce both of these mobile redox carriers (80). Samples with Cu1+ or Zn2+ showed significant changes in the apparent orientations of the gx, gy, and gz axes with respect to the membrane plane (Figure 3E-G and 3H-J). We attempted to simulate the orientation-dependence of the EPR spectra assuming a single orientation of the ISP 2Fe2S cluster. As described under Materials and Methods, we manually superimposed the simulated orientation dependence of the EPR transitions against the experimentally determined orientation dependence, until reasonable fits were obtained. Even the best fits showed small but significant deviations from the data. There are several possible explanations for these deviations, including non-Gaussian mosaic spread (see above) or multiple 2Fe2S cluster orientations. Nevertheless, these simulations qualitatively revealed the effects of metals on ISP conformation. The simulations suggested that for the metal-free control samples, the gx, gy, and gz transition axes of the 2Fe2S cluster were at 41°, 43°, and 18°, respectively, away from the membrane plane, which is consistent with previous results (37, 39). Addition of metal ions shifted all three g axes, to approximately 45°, 37°, and 22° for Cu1+ and to 48°, 32°, and 24° for Zn2+, respectively. The ferricytochrome f EPR transition at gz ) 3.5, which lies perpendicular to the heme plane (81-83), was used to probe the orientation of the cyt f protein with respect to the membrane plane. Figure 4A shows the EPR spectra of the oriented sample in the g ) 3.5 range with the magnetic field oriented at 0° and 60° with respect to the membrane plane, representing the minimal and maximal amplitudes, respectively (see below). Figure 4B is a plot of the orientation dependence of the gz transition of cyt f with respect to the membrane plane. In control samples, the maximum amplitude of the cyt f gz transition was found at 60°, implying that the heme plane lies at 30° from the membrane plane as previously reported in (37, 81, 83). Binding of Cu2+ clearly shifted the gz transition by about 5° to ∼55°, indicating a shift in the orientation of the heme plane from 30° to ∼35° with respect to the membrane plane upon binding of Cu2+. This is consistent with a Cu2+ binding site on the cyt f subunit as suggested in (46, 80). The overlapping gz transitions of both ferrocytochromes bL and bH are found at gz ) 3.6 (81, 83) and were used to probe the orientations of both cyt b hemes. Figure 5A shows EPR spectra in the g ) 3.6 region with the magnetic field at 0° and 90° with respect to the membrane plane. Figure 5B shows the orientation of the gz ) 3.6 signal (37) with respect to the membrane plane. The maximum amplitude of the gz signal was observed at ∼0°, implying that the planes of the cyt b hemes are oriented nearly perpendicular (82) to the membrane plane, as expected from previous EPR results (37) and models of the cyt b6f complex (3, 4, 84). The orientations of the gz transitions were virtually unchanged by Cu1+ with excess ascorbate, indicating that this metal did not perturb

Biochemistry, Vol. 41, No. 12, 2002 4075

FIGURE 4: Effect of Cu2+ on the orientation of the cyt f gz transition. (A) EPR spectra of the cyt f gz transition in partially oriented samples of cyt b6f complex in the absence of metals with magnetic fields at 0° (solid line) or 60° (dashed line) with respect to the membrane plane. (B) Rotational plot of the cyt f heme EPR gz ) 3.5 transition with respect to the membrane plane determined as for the 2Fe2S cluster (Figure 3) in the absence (solid line and closed squares) and presence of Cu2+ (dashed line and open circles). EPR spectrometer settings were the same as in Figure 2, except modulation amplitude ) 3.2 mT; center field ) 210 mT; sweep width ) 200 mT; averages ) 16; microwave power ) 1 mW.

the membrane-spanning domains of the complex, including cyt b6 and subunit IV. These experiments were repeated in the presence of excess ferricyanide, where both copper and cyt f are in their paramagnetic forms (i.e., Cu2+ and ferricytochrome f). In this case, careful analysis of the EPR spectra was required because of overlap with the cyt f gz )3.5 signal. Nevertheless, no apparent effect was seen upon binding of Cu2+ on the orientations of the cyt b hemes (data not shown). Likewise, the addition of Zn2+ had virtually no effect on the orientation of the gz transitions (data not shown). Under conditions used to bind these metals, it is clear that these metals do not significantly disrupt the orientation of the complexes on the Mylar sheets. DISCUSSION Much has been learned from the use of inhibitors which block specific partial reactions of the cyt bc and cyt b6f complexes (reviewed in 1, 85, 86). For example, specific inhibitors first allowed the elucidation of the essential functions of the two Qo site niches and the Qi site (87-91). In addition, such inhibitors have allowed the complex to be poised or trapped in important intermediate states of turnover for detailed spectroscopic study (e.g., 89, 92). Metal ions represent a distinct class of cyt bc and cyt b6f complex inhibitors (41-45). Our work suggested that Cu2+

4076 Biochemistry, Vol. 41, No. 12, 2002

Roberts et al. Scheme 1: Proposed Hypothetical Effects of Metals and Qo Site Inhibitors on the Orientation of the 2Fe2S Cluster and the Soluble Head Domain of the ISPa

FIGURE 5: Effect of metals on the orientation of the cyt b heme gz transition. (A) EPR transitions of the cyt b’s at gz ) 3.6 in the absence of metals with magnetic field at 0° (solid line) and 90° (dashed line) with respect to the membrane plane. (B) Orientational dependence of the gz ) 3.6 transition was determined as described in Figure 4, in the absence (solid line and closed squares) and presence of Cu1+ (dashed line and open circles). EPR spectrometer settings were the same as in Figure 4 except the microwave power was 20 mW.

and Zn2+ inhibit PQH2 binding to the Qo site of the cyt b6f complex by long-range conformational changes (46). We proposed a model wherein ISP pivoting transmitted ‘allosteric’ effects of the metal from their distant binding sites to the Qo site. In essence, we proposed that certain metal ions are specific inhibitors of ISP domain movements, restraining the ISP into distinct positions. Metal Ions Affect the Conformation of the 2Fe2S Cluster of the ‘Rieske” ISP. The data in Figure 3 confirm that metals shift the ISP into distinct positions. To better illustrate these conformational changes, we assigned the directions of the EPR transitions to the specific molecular axis of the 2Fe2S cluster. There is some disagreement in the literature concerning this issue (74, 93-95). Based on pulsed-EPR spectroscopy of the ‘Rieske-type’ 2Fe2S cluster in phthalate dioxygenase, Hoffman and co-workers (93, 94) suggested that the gz transition lies along the Fe-Fe axis. On the other hand, Bertrand et al. (95), using ligand field analysis, suggested that the gx transition occurs along the Fe-Fe axis. The latter assignment is more consistent with conclusions drawn from comparisons of the orientations of the ‘Rieske’ 2Fe2S cluster EPR transitions (37, 39, 96) with the orientation of the ‘Rieske’ 2Fe2S cluster in the X-ray crystal structures of the cyt bc1 complex (e.g., 33). More recent pulsed and oriented EPR results (M. K. Bowman, A. G. Roberts, and D. M. Kramer, in preparation) also suggest assignments in line with the X-ray structures (39), and we tentatively assign gx, gy, and gz to the Fe-Fe, through the plane formed by the 2Fe2S and S-S directions, respectively.

a Ball-and-stick models of the 2Fe2S cluster (top) and solventaccessible surface (bottom) of the ISP (black) and the other subunits of the complex (gray) were generated based on X-ray structures of the cyt bc1 complex (e.g., 33), and oriented based on the data in Figure 3 and the molecular model described in (46). The cyt f protein and the Qo site region are designated f and Qo, respectively. The diagrams show the proposed 2Fe2S cluster orientations and the ISP head domain orientations with respect to the membrane plane, in the (A) absence of inhibitors or metals, (B) presence of copper, (C) presence of Zn2+, and (D) presence of the Qo site inhibitors, DBMIB or stigmatellin. The orientation of the membrane plane is shown with respect to the above diagrams. The arrows in the upper panels show the tentative orientation of the gy EPR transition of the 2Fe2S cluster (see text for details). For comparison, the gray arrows indicate the orientation of the gy axis in the control.

Scheme 1 illustrates the proposed effects of metals, copper and Zn2+, and Qo site inhibitors on the orientation of the 2Fe2S cluster and the hydrophilic head domain of the ISP. Using the above molecular axis assignments, the data in Figures 3 suggest that the addition of Cu1+ or Zn2+ pivoted the ISP roughly around the S-S axis, by about 7° (Scheme 1, B) and about 11° (Scheme 1, C), respectively. Oriented EPR studies of cyt b6f complexes indicated that addition of stigmatellin or DBMIB induced rotation of the ISP around the same (S-S) axis by approximately 60° (as shown in Scheme 1, D), consistent with a shift from the ISPC to ISPB positions (33, 37, 39). Since the complexes were inherently rotationally disordered within the plane of the membrane, it was not possible to determine from our EPR data whether the rotations induced by metal ions and Qo site occupants were in the same direction. However, in the context of the cyt bc1 X-ray structure and models of the cyt b6f complex, the most conservative hypothetical position of the soluble head domain of the ISP would be one which would minimize unfavorable distortions of the neck region of the ISP and would be reasonably close to the Qo site (4, 33, 46, 97). Therefore, we propose that the rotations of the ISP induced by inhibitory metals were in the same general direction as those induced by Qo site occupants. Projecting the approximate rotation data onto the various X-ray crystal structures of the cyt bc1 complex suggests that inhibitory metals lock the ISP into intermediate positions, similar to those observed in some X-ray structures (98). Within the context of our model as described in (46), such

Inhibitory Metals Alter Cytochrome b6f ISP Orientation Scheme 2: Simplified Working Model of the Thermodynamic Relationship between Qo Site Occupants and Metalsa

a The positions of the soluble head domain of the ISP are represented by ISPC, ISPI, ISPI*, and ISPB. A range of other ISP positions is possible (see text), but not shown for clarity. M and Q are metals and Qo site occupants, respectively. The M‚ISPI, Q‚ISPB, and Q‚M‚ISPI* states refer to the positions of the ISP, when metals, Qo site occupants, and both are bound, respectively. A more detailed description of the thermodynamic relationships is given in the text.

constrained intermediate positions would likely reduce the affinity of quinol for the Qo site and would strongly support the role of ISP domain movements in Qo site catalysis. Thermodynamics of Inhibitory Metals and Qo Site Occupants. We propose a thermodynamic relationship between metals and Qo site occupants as diagrammed in Scheme 2. In the absence of metals and distal Qo site occupants, the hydrophilic head domain of the ISP can adopt a range of positions between two extremes, ISPC and ISPB (reviewed in 32, 36). We have designated two distinct intermediate positions between these two extremes, the ISPI and ISPI* positions, which can preferentially bind metal ions. When Qo site occupants, such as quinol or DBMIB, are bound at the distal niche of the Qo site, the ISP appears to be restrained by hydrogen bonding to the ISPB position (33, 37-39). This is designated the Q‚ISPB state in Scheme 2. In the absence of inhibitors under reducing conditions, the cyt bc1 and cyt b6f complex ISP tends to adopt the ISPC position (29, 39). We proposed that bound metals (M), Zn2+, Cu2+ and Cu1+, lock the head domain of the ISP into intermediate positions, which we have collectively designated the M‚ISPI state. By shifting the ISP head domain into the M‚ISPI position, copper ions and Zn2+ would lower the binding affinity of quinol or Qo site inhibitors at the Qo site. Likewise, Qo site occupants should interfere with metal binding by shifting the ISP into the Q‚ISPB state. This is consistent with the competitive behavior of Cu2+ (46) and Zn2+ (see above) observed in kinetic assays with the substrate PQH2. It is also consistent with the apparent displacement of DBMIB from the Qo site by Cu2+ and Zn2+ (Figure 2). Because high concentrations of cyt b6f complex were used in our EPR studies (25-30 µM), we suggest that at least some metals and Qo site occupants are bound to the complexes simultaneously, but in perturbed conformations. Consistent with this, binding of stigmatellin (Figure 1, trace C) and DBMIB (Figure 1, trace B) induced significant

Biochemistry, Vol. 41, No. 12, 2002 4077 distortions in the Cu2+ binding site as deduced by changes in the g| region of the cyt b6f-bound Cu2+ EPR spectra. Both Zn2+ and Cu2+ induced complementary changes in the conformation of the bound Qo site occupant, DBMIB, as shown by shifts in the g-transitions of the DBMIB-shifted 2Fe2S EPR spectrum (Figure 2). Assuming that these metal ions bind at exclusive sites far away from the Qo site as shown for Cu2+ (46), we propose that simultaneous binding of both Qo site occupants and metals shifts the ISP into a distinct position, which we have designated M‚Q‚ISPI*. We have previously concluded that Cu2+ binds to a histidine possibly on cyt f or the ISP, but far from the Qo site itself (46). It is not clear precisely how Cu2+ binding to this residue could shift the ISP. We observed that copper ions shift the orientations of both the ISP and cyt f (Figures 3 and 4). There are two obvious explanations for this. First, the metal ion could directly interact with both subunits, though this appears incompatible with our hypothetical structural model of the cyt b6f complex (46). Second, the metals could bind exclusively on one or the other protein. Since cyt f and the ISP are likely to interact, secondary mutual effects on the orientation may be expected. ACKNOWLEDGMENT We are indebted to Dr. R. Malkin for the donation of his 200tt EPR spectrometer to our laboratory. We thank Drs. W. Nitschke and M. Brugna for valuable discussions and technical assistance in this work. REFERENCES 1. Hauska, G., Schutz, M., and Buttner, M. (1996) in Oxygenic Photosynthesis: The Light Reactions (Ort, D. R., and Yocum, C. F., Eds.) pp 377-398, Kluwer Academic Publishers, Dordrecht, The Netherlands. 2. Berry, E. A., Guergova-Kuras, M., Huang, L.-S., and Crofts, A. R. (2000) Annu. ReV. Biochem. 69, 1005-1075. 3. Cramer, W. A., Martinez, S. E., Huang, D., Tae, G.-S., Everly, R. M., Heymann, J. B., Cheng, R. H., Baker, T. S., and Smith, J. L. (1994) J. Bioenerg. Biomembr. 26, 31-47. 4. Soriano, G. M., Ponamarev, M. V., Carrell, C. J., Xia, D., Smith, J. L., and Cramer, W. A. (1999) J. Bioenerg. Biomembr. 31, 201-213. 5. Hope, A. B. (2000) Biochim. Biophys. Acta 1456, 5-26. 6. Trumpower, B. L., and Gennis, R. B. (1994) Annu. ReV. Biochem. 63, 675-716. 7. Sone, N., Tsuchiya, N., Inoue, M., and Noguchi, S. (1996) J. Biol. Chem. 271, 12457-12462. 8. Yu, J., and Le Brun, N. E. (1998) J. Biol. Chem. 273, 88608866. 9. Kramer, D. M., Schoepp, B., Liebl, U., and Nitschke, W. (1997) Biochemistry 36, 4203-4211. 10. Wollman, F. A., and Lemaire, C. (1988) Biochim. Biophys. Acta 933, 85-94. 11. Allen, J. F., Bennett, J., Steinback, K. E., and Arntzen, C. J. (1981) Nature 291, 25-29. 12. Zito, F., Finazzi, G., Delosme, R., Nitschke, W., Picot, D., and Wollman, F. A. (1999) EMBO J. 18, 2961-2969. 13. Finazzi, G., Zito, F., Barbagallo, R. P., and Wollman, F. A. (2001) J. Biol. Chem. 276, 9770-9774. 14. Kallas, T. (1994) in The Molecular Biology of Cyanobacteria (Bryant, D. A., Ed.) pp 259-317, Kluwer Academic Publishers, Dordrecht, The Netherlands. 15. Martinez, S. E., Huang, D., Szczepaniak, A., Cramer, W. A., and Smith, J. L. (1994) Structure 2, 95-105. 16. Huang, D., Everly, R. M., Cheng, R. H., Heymann, J. B., Schagger, H., Sled, V., Ohnishi, T., Baker, T. S., and Cramer, W. A. (1994) Biochemistry 33, 4401-4409. 17. Pierre, Y., Breyton, C., Lemoine, Y., Robert, B., Vernotte, C., and Popot, J. L. (1997) J. Biol. Chem. 272, 21901-21908.

4078 Biochemistry, Vol. 41, No. 12, 2002 18. Hamel, P., Olive, J., Pierre, Y., Wollman, F.-A., and de Vitry, C. (2000) J. Biol. Chem. 275, 17072-17079. 19. Chain, R. K., and Malkin, R. (1979) Arch. Biochem. Biophys. 197, 52-56. 20. Mitchell, P. (1975) FEBS Lett. 59, 137-139. 21. Crofts, A. R., and Wang, Z. (1989) Photosynth. Res. 22, 6987. 22. Trumpower, B. L. (1990) J. Biol. Chem. 265, 11409-11412. 23. Crofts, A. R. (1985) in The Enzymes of Biological Membranes (Martonosi, A. N., Ed.) pp 347-382, Plenum Publishing, New York. 24. Brandt, U., and Trumpower, B. (1994) Crit. ReV. Biochem. Mol. Biol. 29, 165-197. 25. Brandt, U. (1996) Biochim. Biophys. Acta 1275, 41-46. 26. Snyder, C. H., Gutierrez-Cirlos, E. B., and Trumpower, B. L. (2000) J. Biol. Chem. 275, 13535-13541. 27. Kramer, D. M., and Crofts, A. R. (1993) Biochim. Biophys. Acta 1183, 72-84. 28. Sacksteder, C. A., Kanazawa, A., Jacoby, M. E., and Kramer, D. M. (2000) Proc. Natl. Acad. Sci. U.S.A. 97, 14283-14288. 29. Brandt, U. (1998) Biochim. Biophys. Acta 1365, 261-268. 30. Brandt, U., and von Jagow, G. (1991) Eur. J. Biochem. 195, 163-170. 31. Brandt, U. (1996) FEBS Lett. 387, 1-6. 32. Crofts, A. R., Guergova-Kuras, M., Huang, L., Kuras, R., Zhang, Z., and Berry, E. A. (1999) Biochemistry 38, 1579115806. 33. Zhang, Z., Huang, L., Shulmeister, V. M., Chi, Y. I., Kim, K. K., Hung, L. W., Crofts, A. R., Berry, E. A., and Kim, S. H. (1998) Nature 392, 677-684. 34. Iwata, S., Lee, J. W., Okada, K., Lee, J. K., Iwata, M., Rasmussen, B., Link, T. A., Ramaswamy, S., and Jap, B. K. (1998) Science 281, 64-71. 35. Hunte, C., Koepke, J., Lange, C., Rossmanith, T., and Michel, H. (2000) Struct. Fold. Des. 8, 669-684. 36. Izrailev, S., Crofts, A. R., Berry, E. A., and Schulten, K. (1999) Biophys. J. 77, 1753-1768. 37. Riedel, A., Rutherford, A. W., Hauska, G., Muller, A., and Nitschke, W. (1991) J. Biol. Chem. 266, 17838-17844. 38. Brugna, M., Rodgers, S., Schricker, A., Montoya, G., Kazmeier, M., Nitschke, W., and Sinning, I. (2000) Proc. Natl. Acad. Sci. U.S.A. 97, 2069-2074. 39. Schoepp, B., Brugna, M., Riedel, A., Nitschke, W., and Kramer, D. M. (1999) FEBS Lett. 450, 245-250. 40. Prince, R. C., Crowder, M. S., and Bearden, A. J. (1980) Biochim. Biophys. Acta 592, 323-337. 41. Link, T. A., and von Jagow, G. (1995) J. Biol. Chem. 270, 25001-25006. 42. Skulachev, V. P., Chistyakov, V. V., Jasaitis, A. A., and Smirnova, E. G. (1967) Biochem. Biophys. Res. Commun. 26, 1-6. 43. Nicholls, P., and Malviya, A. N. (1968) Biochemistry 7, 305310. 44. Lorusso, M., Cocco, T., Sardanelli, A. M., Minuto, M., Bonomi, F., and Papa, S. (1991) Eur. J. Biochem. 197, 555561. 45. Kleiner, D. (1974) Arch. Biochem. Biophys. 165, 121-125. 46. Rao, B. K. S., Tyryshkin, A. M., Roberts, A. G., Bowman, M. K., and Kramer, D. M. (1999) Biochemistry 38, 32853296. 47. Berry, E. A., Zhang, Z., Bellamy, H. D., and Huang, L. (2000) Biochim. Biophys. Acta 1459, 440-448. 48. Carrell, C. J., Zhang, H., Cramer, W. A., and Smith, J. L. (1997) Structure 5, 1613-1625. 49. Iwata, S., Saynovits, M., Link, T. A., and Michel, H. (1996) Structure 4, 567-579. 50. Arnon, D. I. (1949) Plant Physiol. 24, 1-15. 51. Renganathan, M., and Bose, S. (1990) Photosynth. Res. 23, 95-99. 52. Hurt, E., and Hauska, G. (1981) Eur. J. Biochem. 117, 591599.

Roberts et al. 53. Hauska, G. (1986) Methods Enzymol. 126, 271-285. 54. Deutscher, M. P., Simon, M. I., and Abelson, J. N. (1997) Guide to Protein Purification, Academic Press, San Diego, CA. 55. Rutherford, A. W., and Setif, P. (1990) Biochim. Biophys. Acta 1019, 128-132. 56. Blasie, J. K., Erecinska, M., Samuels, S., and Leigh, J. S. (1978) Biochim. Biophys. Acta 501, 33-52. 57. More, C., Belle, V., Asso, M., Fournel, A., Roger, G., and Guigliarelli, B. (1999) Biospectroscopy 5, S3-S18. 58. Guigliarelli, B., Guillaussier, J., More, C., Setif, P., Bottin, H., and Bertrand, P. (1993) J. Biol. Chem. 268, 900-908. 59. Poole, C. P., Jr. (1983) Electron Spin Resonance: A ComprehensiVe Treatise on Experimental Techniques, 2nd ed., Dover Publications, Inc., Mineola, NY. 60. Pilbrow, J. R. (1990) Transition Ion Electron Paramagnetic Resonance, 1st ed., Clarendon Press, Oxford, U.K. 61. Vanngard, T. (1972) in Biological Applications of Electron Spin Resonance (Swartz, H. M., Bolton, J. R., and Borg, D. C., Eds.) pp 411-447, Wiley-Interscience, New York. 62. Peisach, J., and Blumberg, W. E. (1974) Arch. Biochem. Biophys. 165, 691-708. 63. Solomon, E. I., Penfield, K. W., and Wilcox, D. E. (1983) Struct. Bond. 53, 1-57. 64. Solomon, E. I., Baldwin, M. J., and Lowery, M. D. (1992) Chem. ReV. 92, 521-542. 65. Weaver, E. C. (1968) Biochim. Biophys. Acta 162, 286-289. 66. Utschig, L. M., Poluektov, O., Schlesselman, S. L., Thurnauer, M. C., and Tiede, D. M. (2001) Biochemistry 40, 6132-6141. 67. Aronoff-Spencer, E., Burns, C. S., Avdievich, N. I., Gerfen, G. J., Peisach, J., Antholine, W. E., Ball, H. L., Cohen, F. E., Prusiner, S. B., and Millhauser, G. L. (2000) Biochemistry 39, 13760-13771. 68. Grogan, J., McKnight, C. J., Troxler, R. F., and Oppenheim, F. G. (2001) FEBS Lett. 491, 76-80. 69. Bertocci, U., and Wagman, D. D. (1985) in Standard Potentials in Aqueous Solution (Bard, A. J., Parsons, R., and Jordan, J., Eds.) pp 287-293, Marcel Dekker, Inc., New York. 70. Huie, R. E., and Neta, P. (1999) in ReactiVe Oxygen Species in Biological Systems: An Interdisciplinary Approach (Gilbert, D. L., and Colton, C. A., Eds.) pp 33-73, Kluwer Academic/ Plenum Publishers, New York. 71. Chevion, M., Berenshtein, E., and Zhu, B.-Z. (1999) in ReactiVe oxygen species in biological systems (Gilbert, D. L., and Colton, C. A., Eds.) pp 103-131, Kluwer Academic/ Plenum Publishers, New York. 72. Brodd, R. J., and Werth, J. (1985) in Standard Potentials in Aqueous Solution (Bard, A. J., and Parsons, R., Eds.) pp 249257, Marcel Dekker, Inc., New York. 73. Trumpower, B. L. (1981) Biochim. Biophys. Acta 639, 129155. 74. Link, T. A., and Iwata, S. (1996) Biochim. Biophys. Acta 1275, 54-60. 75. Malkin, R. (1981) Isr. J. Chem. 21, 301-305. 76. Roberts, A. G., and Kramer, D. M. (2001) Biochemistry 40, 13407-13412. 77. Prince, R. C., Linkletter, S. J. G., and Dutton, P. L. (1981) Biochim. Biophys. Acta 635, 132-148. 78. Malkin, R. (1982) Biochemistry 21, 2945-2950. 79. Baes, C. F., and Mesmer., R. E. (1976) The Hydrolysis of Cations, Wiley, New York. 80. Rao B. K., S. (1998) Mechanism and Location of Inhibitory Copper in b6f Complex: Modes of Copper Inhibition Different in Photosystem I, b6f complex and Photosystem II, Ph.D. Thesis, Washington State University, Pullman, WA. 81. Crowder, M. S., Prince, R. C., and Bearden, A. (1982) FEBS Lett. 144, 204-208. 82. Taylor, C. P. (1977) Biochim. Biophys. Acta 491, 137-48. 83. Bergstro¨m, J., and Vanngard, T. (1982) Biochim. Biophys. Acta 682, 452-456. 84. Widger, W. R., Cramer, W. A., Herrmann, R. G., and Trebst, A. (1984) Proc. Natl. Acad. Sci. U.S.A. 81, 674-678.

Inhibitory Metals Alter Cytochrome b6f ISP Orientation 85. von Jagow, G., and Link, T. A. (1986) Methods Enzymol. 126, 253-271. 86. Link, T. A., Haase, U., Brandt, U., and Jagow, G. v. (1993) J. Bioenerg. Biomembr. 25, 221-232. 87. Ohnishi, T., and Trumpower, B. L. (1980) J. Biol. Chem. 255, 3278-3284. 88. Meinhardt, S. W., and Ohnishi, T. (1992) Biochim. Biophys. Acta 1100, 67-74. 89. Barbagallo, R. P., Finazzi, G., and Forti, G. (1999) Biochemistry 38, 12814-12821. 90. von Jagow, G., and Ohnishi, T. (1985) FEBS Lett. 185, 311315. 91. Brandt, U., Schagger, H., and von Jagow, G. (1988) Eur. J. Biochem. 173, 499-506. 92. Baymann, F., Robertson, D. E., Dutton, P. L., and Mantele, W. (1999) Biochemistry 38, 13188-13199.

Biochemistry, Vol. 41, No. 12, 2002 4079 93. Gurbiel, R. J., Batie, C. J., Sivaraja, M., True, A. E., Fee, J. A., Hoffman, B. M., and Ballou, D. P. (1989) Biochemistry 28, 4861-4871. 94. Gurbiel, R. J., Doan, P. E., Gassner, G. T., Macke, T. J., Case, D. A., Ohnishi, T., Fee, J. A., Ballou, D. P., and Hoffman, B. M. (1996) Biochemistry 35, 7834-7845. 95. Bertrand, P., Guigliarelli, B., Gayda, J.-P., Beardwood, P., and Gibson, J. F. (1985) Biochim. Biophys. Acta 831, 261-266. 96. Schoepp, B., Breton, J., Parot, P., and Vermeglio, A. (2000) J. Biol. Chem. 275, 5284-5290. 97. Mosser, G., Breyton, C., Olofsson, A., Popot, J. L., and Rigaud, J. L. (1997) J. Biol. Chem. 272, 20263-20268. 98. Iwata, M., Bjorkman, J., and Iwata, S. (1999) J. Bioenerg. Biomembr. 31, 169-175. BI015996K